Ab Initio Study of the Carbon-to-Carbon Identity Proton Transfer from

Ab Initio Study of the Carbon-to-Carbon Identity. Proton Transfer from Ketene to Its Anion in the Gas. Phase. Claude F. Bernasconi* and Philip J. Wenz...
0 downloads 8 Views 31KB Size
2430

J. Am. Chem. Soc. 2001, 123, 2430-2431

Ab Initio Study of the Carbon-to-Carbon Identity Proton Transfer from Ketene to Its Anion in the Gas Phase

Table 1. Gas-Phase Acidities (∆H°) and Intrinsic Barriers of Identity Proton Transfers (∆Hq)a MP2/6-311+ G(d,p)// MP2/6-311+ G(d,p)

Claude F. Bernasconi* and Philip J. Wenzel Department of Chemistry and Biochemistry UniVersity of California, Santa Cruz, California 95064 ReceiVed NoVember 2, 2000 ReVised Manuscript ReceiVed December 18, 2000 There is continued and accelerating interest in the factors that determine the Marcus intrinsic barriers and transition state structures (henceforth TS) of proton-transfer reactions.1 Here we report the intrinsic barrier of the identity carbon-to-carbon proton transfer shown in eq 1 as calculated by ab initio methods. We

OdCdCH2 + HCtCsO- h -OsCtH + CH2dCdO (1) also address the question of how the higher degree of unsaturation in ketene compared to acetaldehyde may affect the TS imbalance4 typical for such reactions.5 Additional motivation for this study was provided by results regarding the rate of protonation of ynolate ions such as PhCtCO- by H+ in aqueous solution.6 The kH+ value of 1.34 × 1010 M-1 s-1 is close to the diffusion-controlled limit and much higher than kH+ for the corresponding enolate ion, PhCHdCHO(1.6 × 107 M-1 s-1).7 The higher kH+ value for the ynolate ion could be the result of either a higher pKa or a lower intrinsic barrier. Based on gasphase acidities of ketene and acetaldehyde,8 it is unlikely that PhCtCO- is more basic than PhCHdCO-, making the second explanation more plausible. Our results support this conclusion. We report calculations at the MP2/6-311+G(d,p)//MP2/6311+G(d,p) (MP2//MP2) and B3LYP/6-311+G(d,p) levels; gasphase acidities were also recomputed at the CCSD(T)/6-311+G(2df,2p) level. For details see the Supporting Information.9 Acidities. Our acidities (Table 1) are higher than the ones reported by Smith et al.10 (369.6 kcal/mol at MP4/6-311+G(d,p)) (1) The literature on proton transfers is extensive. Only a few of the most recent references are listed here.2,3 (2) Solution reactions: (a) Nevy, J. B.; Hawkinson, D. C.; Blotny, G.; Yao, X.; Pollack, R. M. J. Am. Chem. Soc. 1997, 119, 12722. (b) Moutier, G.; Peigneux, A.; Vichard, D.; Terrier, F. Organometallics 1998, 17, 4469. (c) Bernasconi, C. F.; Kittredge, K. W. J. Org. Chem. 1998, 63, 1944. (d) Richard, J. P.; William, G.; Gao, J. J. Am. Chem. Soc. 1999, 121, 715. (e) Bernasconi, C. F.; Moreira, J. H.; Huang, L. L.; Kittredge, K. W. J. Am. Chem. Soc. 1999, 121, 1674. (f) Terrier, F.; Moutiers, G.; Pelet, S.; Buncel, E. Eur. J. Org. Chem. 1999, 1771. (3) Theoretical studies: (a) Bernasconi, C. F.; Wenzel, P. J.; Keeffe, J. R.; Gronert, S. J. Am. Chem. Soc. 1997, 119, 4008. (b) Agback, M.; Lunell, S.; Husse´nius, A.; Matsson, O. Acta Chem. Scand. 1998, 52, 541. (c) Beksˇic, D.; Bertra´n, J.; Lluch, J. M.; Hynes, J. T. J. Phys. Chem. A 1998, 102, 3977. (d) Harris, N.; Wei, W.; Saunders, W. H., Jr.; Shaik, S. S. J. Phys. Org. Chem. 1999, 12, 259. (e) Van Verth, J. E.; Saunders, W. H., Jr. Can. J. Chem. 1999, 77, 810. (f) Yamataka, H.; Mustamir; Mishima, M. J. Am. Chem. Soc. 1999, 121, 1023. (g) Lee, I.; Kim, C. K. J. Phys. Org. Chem. 1999, 12, 255. (h) Harris, N.; Wei, W.; Saunders, W. H., Jr.; Shaik, S. S. J. Am. Chem. Soc. 2000, 122, 6754. (i) Bernasconi, C. F.; Wenzel, P. J. J. Org. Chem. 2001, 66, 968. (4) The imbalance refers to the fact that charge delocalization into the π-acceptor lags behind the proton transfer at the transition state. (5) (a) Bernasconi, C. F. Acc. Chem. Res. 1987, 20, 301. (b) Bernasconi, C. F. AdV. Phys. Org. Chem. 1992, 27, 119. (c) Bernasconi, C. F. Acc. Chem. Res. 1992, 25, 9. (6) Chiang, Y.; Kresge, A. J.; Popik, V. V. J. Am. Chem. Soc. 1995, 117, 9165. (7) Keeffe, J. R.; Kresge, A. J. In The Chemistry of Enols; Rappoport, Z., Ed.; Wiley & Sons: New York, 1990; p 399. (8) Lias, S. G.; Bartmess, J. E.; Liebman, J. F.; Levin, R. D.; Holmes, J. L.; Mallard, W. G. Gas-Phase Ion and Neutral Thermochemistry. J. Phys. Chem. Ref. Data 1988, 17, Suppl. 1. (9) See Supporting Information paragraph at the end of this paper.

CH2dCdO CH3CHdOc CH2dCH2 CH3CH3 CH4 c

364.4 367.2 407.9 420.2 418.1

B3LYP/ 6-311+ G(d,p)

∆H° (kcal/mol) 364.0 363.5 406.3 418.4 415.3

CCSD(T)/ 6-311+ G(2df,2p)

exptlb

366.1 367.3 408.4 420.2 418.0

364.9 ( 2.6 365.8 ( 2.9 406.0 ( 2.0 421.0 ( 2.0 416.6 ( 0.8

∆Hq (kcal/mol) CH2dCdO -7.8 (-4.2)d -8.9 (-8.4)d CH3CHdOc -0.3 (2.7)d -1.8 (-1.3)d CH2dCH2 3.9 (7.0)d 3.4 (4.2)d CH3CH3 4.8 (9.3)d 6.0 (7.7)d CH4 c 8.1 (12.1)d 6.7 (7.9)d a ∆Hq ) Hq - H(reactants). b Reference 8. c Reference 3i. d Numbers in parentheses are corrected for BSSE (see text).

but in good agreement with experiment;8 at all levels, except for B3LYP/6-311+G(d,p) ketene is slightly more acidic than acetaldehyde. As shown recently,3i the dominant factor enhancing the acidity of acetaldehyde relative to methane or ethane (Table 1) is resonance stabilization of the enolate ion, worth ca. 36.6 kcal/ mol relative to methane, while the field effect of the CHdO group contributes about 13.3 kcal/mol. The greater acidity of ketene compared to ethene is undoubtedly the result of a similar combination of resonance and field effects of the CdO group. The importance of the resonance effect is apparent from the geometric parameters (Table 2) and the group charges (Table 3). The conversion of ketene to its anion is associated with a significant CdC bond contraction/CdO bond elongation and an opening of the HCC bond angle from 119° to 142°, consistent with a significant contribution of 1b to the resonance hybrid. This

HC h dCdO T HCtC-O1b 1a is similar to the corresponding bond changes for acetaldehyde versus its enolate ion.11 The accumulation of negative charge on the CdO group in the anion (Table 3) further supports the importance of 1b. The fact that the acidity difference between ketene and acetaldehyde is much smaller than that between ethene and ethane implies that the ynolate ion stabilizing effect of CdO is smaller than the enolate ion stabilizing effect of CHdO. The greater ability of the sp2 carbon in ketene compared to the sp3 carbon in acetaldehyde to support the charge apparently reduces the dependence on the substituent. The increased s-character is, of course, also the main reason ethene is more acidic than ethane. Transition State Structure. The changes in geometry and group charges in moving to the TS for the ketene are quite similar to those for the corresponding acetaldehyde reaction (eq 2).11 The

OdCHsCH3 + CH2dCHsO- h -

OsCHdCH2 + CH3sCHdO (2)

largest difference is in the progress of angle deformation (47.1% (10) Smith, B. J.; Radom, L.; Kresge, A. J. J. Am. Chem. Soc. 1989, 111, 8297. (11) (a) Saunders, W. H., Jr. J. Am. Chem. Soc. 1994, 116, 5400. (b) Bernasconi, C. F.; Wenzel, P. J. J. Am. Chem. Soc. 1994, 116, 5405. (c) Bernasconi, C. F.; Wenzel, P. J. J. Am. Chem. Soc. 1996, 118, 10494.

10.1021/ja003858x CCC: $20.00 © 2001 American Chemical Society Published on Web 02/15/2001

Communications to the Editor

J. Am. Chem. Soc., Vol. 123, No. 10, 2001 2431

Table 2. Geometries (MP2//MP2)a parameter

acid

anion

TS

% progress at TS

CH2dCdO rCdC 1.322 1.271 1.293 ∆rCdC -0.051 -0.029 100{|∆rCdC|/rCdC} 3.86 rCdO 1.168 1.229 1.201 ∆rCdO 0.061 0.033 100{∆rCdO/rCdO} 5.22 RHCC 119.1 141.6 129.7 ∆RHCC 22.5 10.6 rC-H b 1.374 rC-C ∆rC-C 100{|∆rC-C|/rC-C} rCdO ∆rCdO 100{∆rCdO/rCdO} Rpyr ∆Rpyr rC-H b

56.9 54.1 47.1

CH3CHdOc 1.503 1.391 1.430 -0.112 -0.073 7.45 1.215 1.271 1.246 0.056 0.031 4.60 51.55 0.00 37.32 -51.55 14.23 1.416

rCdC ∆rCdC rC-H b

1.339

rC-C ∆rC-C rC-H b

1.529

CH2dCH2 1.365 0.026 CH3CH3 1.536 0.007

65.2 55.4

27.6

1.353 0.014 1.424

53.8

1.530 0.001 1.436

a For geometries at the B3LYP level see Table S1.9 the proton in flight. c Reference 3i.

b

H refers to

Table 3. NPA Group Charges (MP2//MP2)a group

acid

anion

diffb

TS

diffc

CH2dCdO CH2(CH) -0.273 -0.660 -0.387 -0.598 -0.325e CdO 0.273 -0.340 -0.613d -0.051 -0.324f H in flight 0.297 CH3CHdO CH3(CH2) -0.021 -0.469 -0.448 -0.384 -0.363e CHdO 0.021 -0.531 -0.522d -0.266 -0.287f H in flight 0.301

n

1.47

1.52

CH2(CH) CH2 H in flight

0 0

CH2dCH2 -0.628 -0.628 -0.400 -0.400e -0.372 -0.372d -0.225 -0.225f (1.06) 0.249

CH3(CH2) CH3 H in flight CH4 H in flight

0 0

CH3CH3 -0.798 -0.798 -0.530 -0.530e -0.202 -0.202d -0.107 -0.107f (0.91) 0.273 0.315

a

Atomic charges as well as charges at the B3LYP level, including Mulliken charges, are reported in Table S2.9 b Anion-acid. c TS-acid. d |difference| ) χ in eq 3. e |difference| ) δ in eq 3. f |difference| ) C δY in eq 3.

for the HCC angle in the ketene reaction versus 27.6% for the pyramidal angle of the acetaldehyde reaction), suggesting that the lag in charge delocalization behind bond proton transfer (imbalance) is less pronounced in the ketene reactions; note that, due to symmetry, the progress of the proton transfer is 50% in both cases. Another measure of imbalance is the n-parameter11b,11c (eq 3), where δY, δC, and χ are group charge differences between TS

n ) log(δY/χ)/log(δC + δY)

(3)

and neutral acid (δY, δC), or between the anion and acid (χ),

respectively (see footnotes d-f in Table 3). The n value of 1.47 for the ketene reaction compares with n ) 1.52 for the acetaldehyde reaction, suggesting the ketene TS is slightly less imbalanced than the acetaldehyde TS, consistent with the conclusion based on the angles RHCC and Rpyr, respectively.13 Barriers. Barriers are reported in Table 1 with and without correction for BSSE by the counterpoise method.14 In view of the controversy as to whether the counterpoise method may lead to over-correction at the MP215 level and the fact that, at a given computational level, the corrections are all very similar for the various reactions, we will focus on the uncorrected values. (1) The barrier for the ketene reaction is much lower than that for the ethene reaction (∆∆Hq ) 11.7 kcal/mol at MP2//MP2, 12.3 kcal/mol at B3LYP). This indicates that stabilization of TS by the CdO group is greater than stabilization of the ynolate ion. This is mainly because each of the two CHdCdO fragments carries more than half a negative charge so that the total substituent effect of the two CdO groups on the TS is greater than the effect of one CdO group on the anion. This situation is analogous to that for eq 2,3i which has a lower barrier than the CH3CH3/CH3CH2- system (∆∆Hq ) 5.1 kcal/mol at MP2//MP2, 7.8 kcal/mol at B3LYP). (2) The barrier for the ketene reaction is significantly lower than that for the acetaldehyde reaction (∆∆Hq ) 7.5 kcal/mol at MP2//MP2, ∆Hq ) 7.1 kcal/mol at B3LYP). This result supports the notion that the higher kH+ value for protonation of PhCtCOcompared to the protonation of PhCHdCHO- is the consequence of a lower intrinsic barrier. (3) A major factor reducing the ketene barrier compared to the acetaldehyde barrier is the greater s-character of the acidic carbon of ketene; it makes the TS tighter (rC-H ) 1.374 versus 1.416 Å) and also allows more effective stabilization by hydrogen bonding than in the acetaldehyde reaction because an sp2 carbon is a better hydrogen bond acceptor than an sp3 carbon.3a,12 This factor should play a role in any comparison between sp2 versus sp3 acids, including ethene versus ethane, although here the difference in ∆Hq is much smaller (0.9 kcal/mol at MP2//MP2; 2.6 kcal/mol at B3LYP). The much larger difference between the ketene and acetaldehyde barriers may be the result of the greater negative charge on the CH groups of the ketene TS (-0.598) compared to the charge on the CH2 groups of the acetaldehyde TS (-0.384); this should increase the hydrogen bonding/ electrostatic stabilization of the ketene TS beyond that arising from the strong s-character. In contrast, the surprisingly small difference in ∆Hq between ethene and ethane may arise from the greater negative charges on the CH2 groups of the ethane TS (-0.530) compared to the charges on the CH groups of the ethene TS (-0.400). The resulting enhanced electrostatic/hydrogen bonding stabilization of the ethane TS could offset some of the inherent advantage of the sp2 hybridization in the ethene reaction. There appears to be some additional TS stabilization of the ethane reaction by the polarizability effect of the methyl group; this notion is supported by the fact that ∆Hq for the CH3CH3/CH3CH2- system is lower than that for the CH4/CH3- system. Acknowledgment is made to the donors of the Petroleum Research Fund, administered by the American Chemical Society, for support of this research (grant PRF 33143-AC4). Supporting Information Available: Tables S1 (energies), S2 (charges), Figure S1 (geometries) and computational details (PDF). This material is available free of charge via the Internet at http://pubs.acs.org. JA003858X (12) (a) Cybulski, S. M.; Scheiner, S. J. Am. Chem. Soc. 1987, 109, 4199. (b) Scheiner, S.; Wang, L. J. Am. Chem. Soc. 1992, 114, 3650. (c) Scheiner, S. J. Mol. Struct. (THEOCHEM) 1994, 307, 65. (13) For the ethene and ethane reactions the n values are close to unity, implying the absence of a significant imbalance. (14) Boys, S. F.; Bernardi, F. Mol. Phys. 1970, 19, 553. (15) (a) Cook, D. B.; Sordo, J. A.; Sordo, T. L. Int. J. Quantum Chem. 1993, 48, 375. (b) Davidson, E. R.; Chakravorty, S. J. Chem. Phys. Lett. 1994, 217, 48.